Quantitative Determination of the Regioselectivity of Nucleophilic

Nov 24, 2008 - Charles P. Casey*, Timothy M. Boller, Joseph S. M. Samec and John R. .... Leslie D. Field , Alison M. Magill , Mohan M. Bhadbhade and S...
1 downloads 0 Views 491KB Size
Organometallics 2009, 28, 123–131

123

Quantitative Determination of the Regioselectivity of Nucleophilic Addition to η3-Propargyl Rhenium Complexes and Direct Observation of an Equilibrium between η3-Propargyl Rhenium Complexes and Rhenacyclobutenes Charles P. Casey,* Timothy M. Boller, Joseph S. M. Samec, and John R. Reinert-Nash Department of Chemistry, UniVersity of Wisconsin-Madison, Madison, Wisconsin 53706 ReceiVed July 31, 2008

PMe3 adds selectively to the central carbon of the η3-propargyl complex [C5Me5(CO)2Re(η3CH2CtCCMe3)][BF4] (1-t-Bu) to form the metallacyclobutene [C5Me5(CO)2Re(CH2C(PMe3)d CCMe3)][BF4] (7). The rate of rearrangement of the metallacyclobutene 7 to η2-alkyne complex [C5Me5(CO)2Re(η2-Me3PCH2CtCCMe3)][BF4] (8) is independent of phosphine concentration, consistent with a dissociative mechanism proceeding via η3-propargyl complex 1-t-Bu. The rate of this rearrangement is 480 times slower than the rate of exchange of PMe3 with the labeled metallacyclobutene 7-d9. This rate ratio provides an indirect measurement of the regioselectivity for addition of PMe3 to the central carbon of η3-propargyl complex 1-t-Bu to give 7 compared to addition to a terminal carbon to give 8. The addition of PPh3 to 1-t-Bu gives the metallacyclobutene [C5Me5(CO)2Re(CH2C(PPh3)dCCMe3)][BF4] (11). Low-temperature 1H NMR spectra provide evidence for an equilibrium between metallacyclobutene 11 and η3-propargyl complex 1-t-Bu (Keq ≈ 44 M-1 at -46 °C and ∆G°(0 °C) ) -1.2 ( 0.2 kcal mol-1). Introduction η3-Propargyl transition metal complexes, the triple-bond analogues of the very well studied η3-allyl complexes, are becoming increasingly important in synthetic chemistry.1 The first stable η3-propargyl complex, [(Me3P)4Os(η3-PhCt CCdCHPh)][PF6], was reported by Werner in 1985,2 and the first isolated η3-propargyl complex not having an exo-double bond, [(C6Me5H)Mo(CO)2(η3-CH2CtCH)][BF4], was reported by Kryvich in 1991.3

We first synthesized η3-propargyl rhenium complexes by hydride abstraction from alkyne complexes and later developed a regioselective synthesis via protonation of propargyl alcohol complexes (Scheme 1).4 Other methods for the synthesis of isolable η3-propargyl complexes include halide abstraction from * Corresponding author. E-mail: [email protected]. (1) Reviews: (a) Chen, J.-T. Coord. Chem. ReV. 1999, 190-192, 1143. (b) Kurosawa, H.; Ogoshi, S. Bull. Chem. Soc. Jpn. 1998, 71, 973. (c) Doherty, S.; Corrigan, J. F.; Carty, A. J.; Sappa, E. AdV. Organomet. Chem. 1995, 37, 39. (d) Wojcicki, A. New J. Chem. 1994, 18, 61. (2) Gotzig, J.; Otto, H.; Werner, H. J. Organomet. Chem. 1985, 287, 247. (3) (a) Krivykh, V. V.; Taits, E. S.; Petrovskii, P. V.; Struchov, Y. T.; Yanovski, A. I. MendeleeV Commun. 1991, 103. (b) Gusev, O. V.; Krivykh, V. V.; Rubezhov, A. Z. Organomet. Chem. USSR 1991, 4, 233. (4) (a) Casey, C. P.; Yi, C. S. J. Am. Chem. Soc. 1992, 114, 6597. (b) Casey, C. P.; Selmeczy, A. D.; Nash, J. R.; Yi, C. S.; Powell, D. R.; Hayashi, R. K. J. Am. Chem. Soc. 1996, 118, 6698.

Scheme 1

η1-propargyl or η1-allenyl metal halide complexes,5 reaction of metal halides with propargyl nucleophiles,6 reaction of propargyl ether complexes with Lewis acids,7 rearrangement of η1homopropargyl metal complexes,8 and oxidative addition of propargyl halides or tosylates to metal complexes.6,9 1 H and 13C NMR spectroscopic data and X-ray crystallographic structures suggest that η3-propargyl complexes are best represented as a combination of η3-propargyl and η3-allenyl resonance structures (Figure 1).1 Crystallographic data show that all three carbons are within bonding distance of the metal (2.2-2.5 Å), with the central carbon often nearest to the metal center. The CR2dC bond length is substantially longer (1.34-1.40 (5) (a) Blosser, P. W.; Schimpff, D. G.; Gallucci, J. C.; Wojcicki, A. Organometallics 1993, 12, 1993. (b) Huang, T.-M.; Chen, J.-T.; Lee, G.H.; Wang, Y. J. Am. Chem. Soc. 1993, 115, 1170. (c) Huang, T.-M.; Hsu, R.-H.; Yang, C.-S.; Chen, J.-T.; Lee, G.-H.; Wang, Y. Organometallics 1994, 13, 3657. (d) Ogoshi, S.; Tsutsumi, K.; Kurosawa, H. J. Organomet. Chem. 1995, 493, C19. (e) Baize, M. W.; Blosser, P. W.; Plantevin, V.; Schimpff, D. G.; Gallucci, J. C.; Wojcicki, A. Organometallics 1996, 15, 164. (6) (a) Blosser, P. W.; Gallucci, J. C.; Wojcicki, A. J. Am. Chem. Soc. 1993, 115, 2994. (b) Gordon, G. J.; Whitby, R. J. J. Chem. Soc., Chem. Commun. 1997, 1045. (7) Weng, W.; Arif, A. M.; Gladysz, J. A. Angew. Chem., Int. Ed. Engl. 1993, 32, 891. (8) Stang, P. J.; Crittell, C. M.; Arif, A. M. Organometallics 1993, 12, 4799. (9) Nishida, T.; Ogoshi, S.; Tsutsumi, K.; Fukunishi, Y.; Kurosawa, H. Organometallics 2000, 19, 4488.

10.1021/om800739j CCC: $40.75  2009 American Chemical Society Publication on Web 11/24/2008

124 Organometallics, Vol. 28, No. 1, 2009

Casey et al. Scheme 3

Figure 1. Resonance structures for η3-propargyl complexes.

Å) than in free allenes, and the CtCR bond length is longer than in free acetylenes. The CR2-CtCR angle varies from 145° to 155°, implying substantial strain about the central carbon. This strain may be responsible for the highly reactive nature of η3-propargyl complexes. Scheme 2

carbon to give an allene complex was prevented by steric hindrance (Scheme 3). In some cases, nucleophilic additions to η3-propargyl complexes have led to η3-allyl complexes, consistent with initial kinetic attack at the center carbon of the η3-propargyl ligand followed by protonation of the metallacyclobutene intermediate.12,6 In other cases, nucleophilic additions to η3-propargyl complexes have led to η3-trimethylenemethane complexes, consistent with initial kinetic attack at the center carbon of the η3-propargyl ligand followed by protonation of the metallacyclobutene intermediate and deprotonation at the atom R to the central allyl carbon (Scheme 4).13 Scheme 4

In contrast to η3-allyl metal complexes, which are usually attacked by nucleophiles at a terminal carbon, η3-propargyl metal complexes are selectively attacked by nucleophiles at the center carbon.1 Nucleophiles have been observed to attack each of the three carbon centers of η3-propargyl rhenium complexes. We found that while kinetic addition of nucleophiles occurs at the center carbon of the η3-propargyl ligand to give rhenacyclobutene complexes, this addition is sometimes reversible and can be followed by addition at either of the other two metalbound carbon atoms to give η2-allene or η2-alkyne complexes.4a,10,11 The methyl-substituted η3-propargyl complex [C5Me5(CO)2Re(η3CH2CtCMe)][BF4] (1-Me) reacted irreversibly with carbon nucleophiles such as malonates and acetylides to give metallacyclobutenes (Scheme 2); protonation of the metallacyclobutenes led to η3-allyl complexes. At low temperature, nitrogen nucleophiles added to the central propargyl carbon to produce the metastable metallacyclobutene complexes. Upon warming, the nitrogen nucleophile dissociated from the rhenacyclobutenes and then readded to the resulting η3-propargyl complex at one of the terminal sites to give either an η2-allene or an η2-alkyne complex. For example, 1-Me reacted with pyridine at -40 °C to give rhenacyclobutene 3, which rearranged to η2-allene complex 4 at room temperature (Scheme 2). 4-Dimethylaminopyridine added to the central carbon of the tert-butyl-substituted 1-t-Bu below -40 °C to give the rhenacyclobutene complex 5, which rearranged to η2-alkyne complex 6 at room temperature; apparently, attack at the t-Bu-substituted (10) Casey, C. P.; Nash, J. R.; Yi, C. S.; Selmeczy, A. D.; Chung, S.; Powell, D. R.; Hayashi, R. K. J. Am. Chem. Soc. 1998, 120, 722. (11) The Pd-catalyzed reaction of propargyl halides with thiols gives mixtures of propargyl and allenyl sulfides. These reactions have been proposed to occur by attack of nucleophiles at the terminal carbons of an η3-propargyl intermediate. However, kinetic attack at the central propargyl carbon might be reversible. Tsutsumi, K.; Yabukami, T.; Fujimoto, K.; Kawase, T.; Morimoton, T.; Kakiuchi, K. Organometallics 2003, 22, 2996.

Tsuji pioneered the development of palladium-catalyzed reactions for addition of two nucleophiles to propargyl substrates.14 For example, the reactions of carbon nucleophiles with propargyl carbonates produce double nucleophilic addition products where nucleophiles have added to both the central and a terminal carbon of the propargyl unit. We have proposed that these double nucleophilic additions occur by oxidative addition to produce an η3-propargyl intermediate, which then undergoes addition of the first nucleophile to the central propargyl carbon to produce a metallacyclobutene; this is followed by protonation of the metallacyclobutene to generate an η3-allyl complex that is attacked by a second nucleophile (Scheme 5).10 Scheme 5

Ohno has imaginatively constructed bicyclic heterocycles by palladium-catalyzed domino cyclization of propargyl bromides.15 Yoshida has devised a clever palladium-catalyzed threecomponent coupling of propargylic oxiranes, phenols, and

Nucleophilic Addition to η3-Propargyl Re Complexes Scheme 6

Organometallics, Vol. 28, No. 1, 2009 125

Indirect Determination of Regioselectivity of Attack of Phosphines. The relative rates of nucleophilic addition to the center and terminal carbons can be measured indirectly by comparing the rates of exchange of nucleophiles with the metallacyclobutene adducts and of rearrangement of the metallacyclobutene to an η2-alkyne complex (eqs 1-4, Scheme 7, and Figure 2). The difference between the barriers for exchange and rearrangement is also the difference between the barriers for attack at the central and terminal carbons. This method assumes a dissociative mechanism for nucleophilic exchange with the metallacyclobutene that proceeds via an η3-propargyl intermediate. Scheme 7

carbon dioxide.16 Both of these processes involve double nucleophilic addition to propargyl substrates (Scheme 6).17 Because rhenacyclobutenes are the only observed kinetic product of nucleophilic attack on η3-propargyl rhenium complexes, it was not possible to quantitatively measure the regioselectivity from product ratios. Here we report an indirect method for the quantitative determination of this regioselectivity. We also report the first direct observation of an equilibrium between an η3-propargyl metal complex and a metallacyclobutene.

Results Both the rate and the regioselectivity of the addition of nucleophiles to the central carbon of η3-propargyl rhenium complexes are so high that it is not possible to directly measure how much faster attack at the center carbon is than attack at one of the terminal carbons. Since the rhenacyclobutene is the only observed initial product and neither an η2-alkyne complex nor an η2-allene complex is initially seen, estimates of >20 can be placed on the regioselectivity. In some cases, the initially formed metallacyclobutene rearranged to a thermodynamically more stable η2-alkyne complex (∆G° e -2.3 kcal mol-1). The reaction was complete before the cold sample was inserted into the precooled NMR probe; this requires a t1/2 of less than a minute at -50 °C and ∆Gq(0.02 M ligand) e 14.7 kcal mol-1, ∆Gq(1.0 M ligand) e 12.9 kcal mol-1. (12) Tsai, F.-Y.; Hsu, R.-H.; Huang, T.-M.; Chen, J.-T.; Lee, G.-H.; Wang, Y. J. Organomet. Chem. 1996, 520, 85. (13) (a) Baize, M. W.; Furilla, J. L.; Wojcicki, A. Inorg. Chim. Acta 1994, 223, 1. (b) Baize, M. W.; Plantevin, V.; Gallucci, J. C.; Wojcicki, A. Inorg. Chim. Acta 1995, 235, 1. (c) Su, C.-C.; Chen, J.-T.; Lee, G.-H.; Wang, Y. J. Am. Chem. Soc. 1994, 116, 4999. (d) Plantevin, V.; Blosser, P. W.; Gallucci, J. C.; Wojcicki, A. Organometallics 1994, 13, 3651. (14) (a) For reviews, see: Tsuji, J.; Mandai, T. Angew. Chem., Int. Ed. Engl. 1995, 34, 2589. (b) Ohno, H. Chem. Pharm. Bull. 2005, 53, 1211. (c) Labrosse, J.-R.; Lhoste, P.; Sinou, D. J. Org. Chem. 2001, 66, 6634. (15) Ohno, H.; Okanao, A.; Kosaka, S.; Tsukamoto, K.; Ohata, M.; Ishihara, K.; Maeda, H.; Tanaka, T.; Fujii, N. Org. Lett. 2008, 10, 1171. (16) Yoshida, M.; Murao, T.; Sugimoto, K.; Ihara, M. Synlett 2007, 575. (17) Another interesting synthetic application that involves propargyl complexesisthepalladium-catalyzeddeoxygenationofpropargylformates.Radinov, R.; Hutchings, S. D. Tetrahedron Lett. 1999, 40, 8955.

Rate(exchange) ) k-1[7]

(1)

Rate(rearrangement) ) k-1[7](k2[PMe3] ⁄ {k1[PMe3] + k2[PMe3]} ) k-1[7](k2 ⁄ {k1 + k2}) (2) Rate(rearrangement) ⁄ Rate(exchange) ) k2 ⁄ {k1 + k2} (3) Rate(rearrangement) ⁄ Rate(exchange) ≈ k2 ⁄ k1 when k1 . k2 (4) We chose to study the exchange and rearrangement kinetics of [C5Me5(CO)2Re(CH2C(PMe3)dCCMe3)][BF4] (7) since this metallacyclobutene is relatively stable and had been fully characterized by X-ray crystallography.10 In addition, 7 had been found to slowly rearrange at room temperature to a single product, the η2-alkyne complex [C5Me5(CO)2Re(η2-Me3-

Figure 2. Free energy diagram for nucleophilic addition to η3-propargyl complex 1-t-Bu.

126 Organometallics, Vol. 28, No. 1, 2009

Figure 3. X-ray structure of [C5Me5(CO)2Re(η2-Me3PCH2Ct CCMe3)]+.

PCH2CtCCMe3)][BF4] (8),10 which was further characterized in the current work by X-ray diffraction (Figure 3). The rearrangement product, η2-alkyne complex 8, results from nucleophilic addition to the less crowded terminal CH2 carbon of the intermediate η3-propargyl complex 1-t-Bu and not to the more sterically hindered tert-butyl-substituted carbon to form an η2-allene isomer. Since our method for determining relative regioselectivity depends on the assumption that rearrangement of metallacyclobutene 7 to η2-alkyne complex 8 proceeds by phosphine dissociation to generate an η3-propargyl intermediate, we needed to establish the rate law for the rearrangement. This was particularly necessary since associative mechanisms are conceivable. For example, phosphine attack might occur at the terminal carbon of 7 to give intermediate A, which could then undergo loss of phosphine to generate 8 (Scheme 8). Alkyl and vinyl CpRe(CO)2R- anions are well-known stable species.18 The dissociative mechanism of Scheme 7 and the associative mechanism of Scheme 8 are readily distinguished by their kinetic independence or dependence on [PR3]. Scheme 8

The rate of rearrangement of metallacyclobutene complex 7 was studied in the presence and absence of added PMe3. Two yellow CD3CN solutions19 of 7 were prepared in resealable NMR tubes, and excess PMe3 was condensed into one. The rate of rearrangement of 7 to 8 was monitored by 1H NMR spectroscopy at 64 °C. First-order rate law fits were obtained. The rate in the absence of added PMe3 (krearr ) (8.2 ( 0.1) × 10-4 s-1) and in the presence of 0.29 M added PMe3 (krearr ) (7.9 ( 0.1) × 10-4 s-1) were very similar. The failure to observe a rate enhancement in the presence of added PMe3 rules out an associative mechanism for the rearrangement of metallacyclobutene complex 7 to η2-alkyne complex 8.20 (18) (a) Casey, C. P.; Vosejpka, P. C.; Askham, F. R. J. Am. Chem. Soc. 1990, 112, 3713. (b) Casey, C. P.; Vosejpka, P. C.; Gavney, J. A., Jr. J. Am. Chem. Soc. 1990, 112, 4083. (19) Phosphine reactions in CD2Cl2 were problematic at elevated temperatures due to nucleophilic attack of PMe3 on the solvent. (20) The intriguing possibility of an associative mechanism for rearrangement of a metallacyclobutene was explored by studying rearrangements involving 1,2-bis(diphenylphosphino)ethane ligands, where the associative mechanism would be more likely with the possibility of intramolecular attack by a second phosphine. Even with diphos ligands, the dissociative mechanism is strongly favored. See Supporting Information.

Casey et al.

The deuterium-labeled metallacyclobutene [C5Me5(CO)2Re(CH2C[P(CD3)3]dCCMe3)][BF4] (7-d9) was prepared by addition of P(CD3)3 to 1-t-Bu. The kinetics of the exchange of a 4.5-fold excess of P(CH3)3 with 0.1 M 7-d9 were studied by 1H NMR spectroscopy at 10 °C in CD3CN by monitoring the appearance of the P(CH3)3 resonance of the exchanged metallacyclobutene 7 (Figure 4). Data analysis gave kexch ) (4.9 ( 0.5) × 10-5 s-1 and ∆Gq10 °C ) 22.4 kcal · mol-1. Since exchange of phosphine with 7 is much faster than rearrangement to 8, it was not possible to measure both rates at the same temperature. We therefore measured the rate of rearrangement of 7 to 8 by 1H NMR spectroscopy in CD3CN between 35 and 65 °C. A linear Eyring plot of ln[k/T] versus 1/T gave ∆Hq ) 30.1 ( 0.9 kcal · mol-1, ∆Sq ) 16.0 ( 0.9 eu, and ∆Gq10 °C ) 25.6 kcal · mol-1. Extrapolation of the rearrangement rate to 10 °C gave krearr ) 1.02 × 10-7 s-1. The rate of exchange of PMe3 with metallacyclobutene 7 is therefore 480 times faster than the rearrangement of 7 to 8 at 10 °C (kexch/krearr ) [4.9 × 10-7 s-1] ÷ [1.0 × 10-7 s-1] ) 480). Inspection of Figure 2 reveals that this is also the rate difference between attack of PMe3 at the center carbon and the CH2 terminal carbon of η3-propargyl complex 1-t-Bu. This rate difference corresponds to a 3.2 kcal · mol-1 difference in free energy of activation (∆∆Gq10 °C ) 25.6 - 22.4 kcal · mol-1). Regioselectivity of addition of PMePh2 to 1-t-Bu was also determined indirectly by comparing the rates of rearrangement and exchange of rhenacyclobutene [C5Me5(CO)2Re(CH2C[PMePh2]dCCMe3)][BF4] (9). Addition of PMePh2 to η3propargyl complex 1-t-Bu at -88 °C led to complete conversion to the rhenacyclobutene 9, which had characteristic 1H NMR resonances at δ 0.11 (d, J ) 12.6 Hz, CHH), 1.49 (d, J ) 12.2 Hz, CHH), and 2.39 (d, JPH ) 12.2 Hz, PCH3). The reaction of excess PMe3 with PPh2Me-substituted metallacycle 9 led to exchange and formation of the more stable PMe3-substituted metallacycle 7. The rate constant for exchange at -40 °C was determined by 1H NMR spectroscopy: k ) 5.85 × 10-4 s-1, ∆Gq ) 16.9 kcal · mol-1. Upon warming to room temperature, 9 rearranged to η2-alkyne complex [C5Me5(CO)2Re(η2-Ph2MePCH2Ct CCMe3)][BF4] (10), which had characteristic 1H NMR resonances at δ 2.45 (d, JPH ) 13.1 Hz, PCH3), 4.15 (t, J ) 16.1 Hz, CHH), and 4.72 (t, J ) 15.8 Hz, CHH). The rate of rearrangement of 9 to 10 at -15 °C was measured by 1H NMR spectroscopy: k ) 3.27 × 10-5 s-1, ∆Gq(-15 °C) ) 20.3 kcal · mol-1. The difference between this barrier for rearrangement and the barrier for exchange (∆∆Gq ) 20.3 - 16.9 ) 3.4 kcal · mol-1) is similar to the difference in barriers for exchange and rearrangement of PMe3-substituted metallacycle 7 and implies a regioselective preference of 960 for addition of PMePh2 to the center propargyl carbon of 1-t-Bu. More Rapid Dissociation of PPh3 than PMe3 from Metallacyclobutenes. Previously, we had reported that the addition of PPh3 to 1-t-Bu gave the metallacyclobutene complex [C5Me5(CO)2Re(CH2C(PPh3)dCCMe3)][BF4] (11), which subsequently rearranged to alkyne complex [C5Me5(CO)2Re(η2Ph3PCH2CtCCMe3)][BF4] (12). In repeating this work, we monitored the reaction of 1-t-Bu (0.030 M) with PPh3 (0.063 M) in CD2Cl2 by 1H NMR spectroscopy and found that metallacycle 11 was formed at -78 °C with characteristic lowfrequency doublets for the metal-bound CH2 group at δ 0.14 and 1.1 (2J ) 12 Hz). When the sample was warmed to -20 °C, the conversion of metallacycle 11 to alkyne complex 12 was observed by 1H NMR spectroscopy to be 80% complete within 2 h (∆Gq ≈ 18.9(5) kcal · mol-1). Upon warming to 25

Nucleophilic Addition to η3-Propargyl Re Complexes

Organometallics, Vol. 28, No. 1, 2009 127

Figure 4. Elapsed time 1H NMR spectra for exchange at 10 °C in CD3CN: 7-d9 + P(CH3)3 approaches equilibrium with 7 + P(CD3)3. Peak 4, P(CH3)3 of 7. Other peaks: peak 2, Cp* of 7; 5, CH of 7; 6, tBu of 7; 7, excess P(CH3)3; 1, toluene standard; 3, CD3CN. Scheme 9

°C, only η2-alkyne complex 12 was observed by 1H NMR spectroscopy (Scheme 9). Note that this rate of rearrangement conflicts with the preViously reported slow rearrangement of the metallacycle 11 to η2-alkyne complex 12 oVer 2 days at room temperature in CD2Cl2.10 The much more rapid rearrangement of the PPh3-substituted rhenacyclobutene 9 than the PMe3-substituted rhenacyclobutene 7 (∆∆Gq ≈ 6.7(5) kcal · mol-1) implies a correspondingly more rapid dissociation of phosphine from their respective rhenacyclobutenes and can be attributed in part to the greater leaving group ability of the much less nucleophilic PPh3 group. In addition, the particularly unfavorable steric interaction between the large PPh3 and the bulky tert-butyl group is relieved upon PPh3 dissociation. The much less congested methyl-substituted

metallacyclobutene [C5Me5(CO)2Re(CH2C(PPh3)dCCH3)][BF4] does not rearrange to either an η2-alkyne or η2-allene complex even upon extended heating at 50-60 °C. Only eventual decomposition is observed. Observation of Equilibrium between η3-Propargyl Complex 1-t-Bu and PPh3-Substituted Metallacyclobutene 11. Upon close examination of the low-temperature 1H NMR spectra, it became apparent that the formation of metallacyclobutene 11 did not go to completion and that an equilibrium was established with the η3-propargyl complex 1-t-Bu. After PPh3 (0.063 M in CD2Cl2) was added to a yellow solution of η3-propargyl complex 1-t-Bu (0.030 M in CD2Cl2) at -78 °C, 1 H NMR spectroscopy at -88 °C showed a tiny resonance at δ 3.41 (CHH) characteristic of 1-t-Bu in addition to intense resonances for the metallacyclobutene 11 (δ 0.14, CHH). Upon increasing the temperature to -66 °C, the amount of η3propargyl complex 1-t-Bu increased and a 9.6:1 equilibrium mixture of 11/1-t-Bu was observed (Keq ) [11]/[ 1-t-Bu][PPh3] ≈ 280 M-1). At -56 °C, the amount of the η3-propargyl complex increased further, and a 4.5:1 equilibrium mixture of 11/1-t-Bu was observed (Keq ≈ 100 M-1). When the sample was recooled to -66 °C, the ratio of 11: 1-t-Bu reverted to 9.6:1. At -46 °C, the ratio of 11:1-t-Bu dropped to 1.9:1 (Keq ≈ 44 M-1). When the temperature was lowered to -66 °C, the ratio of 11:1-t-Bu increased to 10.8:1 (Keq ≈ 310 M-1). Thus, as the temperature increases, the equilibrium shifts toward η3propargyl complex 1-t-Bu and free PPh3 at the expense of the adduct rhenacyclobutene 11 (Figure 5). When the sample was warmed to -34 °C, the ratio of 11: 1-t-Bu fell to 1.1:1 (Keq ≈ 25 M-1) and some conversion to η2-alkyne complex [C5Me5(CO)2Re(η2-Ph3PCH2Ct CCMe3)][BF4] (12) (16% after 10 min) was seen (∆Gq ≈ 17.5(5) kcal · mol-1).21 At -23.5 °C, a 1:2 ratio of 11:1-t-Bu (Keq ≈ 11 M-1) was seen along with 56% conversion to the thermodynamically stable η2-alkyne complex 12 after 10 min (∆Gq ≈

128 Organometallics, Vol. 28, No. 1, 2009

Casey et al.

Figure 5. Ratio of rhenacyclobutene 11 to η3-propargyl complex 1-t-Bu as temperature is raised and lowered.

17.8(5) kcal mol-1).21 After 1 h at room temperature, complete conversion to 12 was seen. When this solution of η2-alkyne complex 12 was recooled to -66 °C, no change in the NMR spectrum was seen, demonstrating that the formation of the thermodynamically stable η2-alkyne complex is irreversible. Additional measurements of the equilibrium between the rhenacyclobutene and the η3-propargyl complex carried out using lower concentrations of the PPh3 (0.042 M in CD2Cl2) and the same concentration of η3-propargyl complex 1-t-Bu (0.03 M in CD2Cl2) gave similar values for the equilibrium constants. At -66 °C, a 4.5:1 equilibrium mixture of rhenacyclobutene 11:η3-propargyl complex 1-t-Bu was observed by 1H NMR spectroscopy (Keq ≈ 260 M-1). At -46 °C, a 1:1 mixture of 11/1-t-Bu was seen (Keq ≈ 35 M-1). A van’t Hoff plot of all the equilibrium data gave ∆H° ) -8.0 ( 1 kcal · mol-1, ∆S° ) -25 ( 3 eu, and ∆G°(0 °C) ) -1.2 ( 0.2 kcal · mol-1. Equilibrium between η3-Propargyl Complex 1-t-Bu and Other Phosphine-Substituted Metallacyclobutenes. The more nucleophilic phosphine PMePh2 formed the more stable metallacyclobutene 9, and the less nucleophilic phosphine P(C6H4p-F)3 formed the less stable metallacyclobutene 13. Addition of PMePh2 (0.05 M in CD2Cl2) to η3-propargyl complex 1-t-Bu (0.030 M in CD2Cl2) at -88 °C led to complete conversion to rhenacyclobutene 9. Neither η3-propargyl complex 1-t-Bu nor the stable η2-alkyne complex 10 was observed by 1 H NMR spectroscopy below -24 °C. However, upon warming to -13 °C, 1H NMR spectroscopy showed small amounts of the η3-propargyl complex 1-t-Bu (8%) and of η2-alkyne complex 10 (17% after 10 min and increasing amounts at longer times). When the temperature was raised to 8 °C, the amount of η3propargyl complex 1-t-Bu remained relatively constant (8%) while the amount of rhenacyclobutene complex 9 decreased and the amount of η2-alkyne complex 10 increased. At lower than stoichiometric concentrations of PMePh2 (0.011 M in CD2Cl2) compared to 1-t-Bu (0.030 M in CD2Cl2), a relatively constant 0.43:1 ratio of rhenacyclobutene complex 9 to η3-propargyl complex 1-t-Bu was seen between -88 and -56 °C. Too little free PMePh2 was present to be observed. When the temperature was increased to -24 °C, some dissociation of PMePh2 from the metallacycle occurred and the ratio of 9:1-tBu decreased to 0.32:1 (Keq ≈ 80 M-1, ∆G(-24 °C) ≈ -2.2 (21) These approximate ∆G values are obtained from the approximate rates. If rate estimates are off by a factor of 3 at 0 °C, ∆Gq would be off by 0.6 kcal · mol-1. q

kcal · mol-1).22 When the solution was recooled to -66 °C, the ratio reverted to 0.43:1. When the solution was warmed to -2 °C, the ratio dropped to 0.2:1 (Keq ≈ 50 M-1, ∆G°(-2 °C) ≈ -2.1 kcal · mol-1) and some rearrangement to the thermodynamically stable η2-alkyne complex 10 was observed (6% after 5 min). Upon further increasing the temperature to 9 °C, the ratio of 9:1-t-Bu dropped to 0.1:1 (Keq ≈ 25 M-1, ∆G°(9 °C) ≈ -1.8 kcal · mol-1). After 15 min at 9 °C, all the PMePh2 was consumed and only propargyl complex 1-t-Bu and η2alkyne complex 10 were observed. A plot of ∆G versus temperature allowed extrapolation to ∆G°(0 °C) ≈ -2.0 kcal · mol-1.

Figure 6. Thermodynamics of rhenacyclobutene formation.

Addition of the less nucleophilic phosphine P(C6H4-p-F)3 (0.05 M in CD2Cl2) to η3-propargyl complex 1-t-Bu (0.030 M in CD2Cl2) at -80 °C led to slow conversion (∆Gq (metallacycle formation) ≈ 14.2(8) kcal mol-1)21 to the rhenacyclobutene [C5Me5(CO)2Re(CH2C[P(C6H4-p-F)3]dCCMe3)][BF4] (13), which had characteristic 1H NMR resonances at δ 0.05 (CHH) and 0.70 (CMe3). After 2 h, a 1.3:1 equilibrium mixture of 13/1-tBu was observed (Keq ) [13]/[1-t-Bu][PAr3] ≈ 38 M-1). When the solution was warmed to -70 °C, the ratio dropped to 0.56:1 (Keq ≈ 14 M-1). Upon further increasing the temperature to -60 °C, the ratio of 13:1-t-Bu dropped to 0.2:1 (Keq ≈ 4.6 M-1). Conversion to the η2-alkyne complex [C5Me5(CO)2Re[η2-(p-FC6H4)3PCH2CtCCMe3][BF4] (14) began at -30 °C and was complete at -20 °C (∆Gq ≈ 17.2(8) kcal · mol-1).21 (22) Because the free phosphine was difficult to observe directly, its concentration was estimated by subtracting the concentration of rhenacyclobutene complex and η2-alkyne complex from the initial phosphine concentration. Errors in estimation are considerable, and only very approximate equilibrium constants were obtained.

Nucleophilic Addition to η3-Propargyl Re Complexes

Organometallics, Vol. 28, No. 1, 2009 129

Additional measurements of the equilibrium between the 13 and 1-t-Bu were carried out using a higher concentration of the P(C6H4-p-F)3 (0.10 M in CD2Cl2), and the same concentration of 1-t-Bu (0.03 M in CD2Cl2) gave similar values for the equilibrium constants. [At -80 °C, 4.3:1 13:1-t-Bu, Keq ≈ 54 M-1; at -70 °C, 1.7:1 13:1-t-Bu, Keq ≈ 20 M-1; at -60 °C, 0.7:1 13:1-t-Bu, Keq ≈ 7.6 M-1.] A van’t Hoff plot gave ∆H° ) -7.2 ( 0.5 kcal · mol-1, ∆S° ) -30 eu ( 4, and ∆G°(0 °C) ) -1.0 ( 0.2 kcal · mol-1. The equilibrium between PMe3-substituted metallacycle 7 and η3-propargyl complex 1-t-Bu lay too far on the side of the metallacycle to be directly observable. A limit on the equilibrium constant can be estimated from the fact that none of 1-t-Bu was detectable by 1H NMR spectroscopy of 0.1 M solutions of 7, and 2% of 1-t-Bu would have been readily detected. This requires Keq g 25 000 M-1 and ∆G°(0 °C) e -5.5 kcal · mol-1. Figure 5 summarizes the thermodynamic information of the equilibrium between rhenacyclobutenes and η3-propargyl complex 1-t-Bu.

Discussion Thermodynamics of Rhenacyclobutene Formation. It is surprising that the metallacyclobutene complexes and η3-propargyl complex 1-t-Bu are so similar in thermodynamic stability. For PPh3, PPh2Me, and P(C6H4-p-F)3, both species were observable in solution and Keq could be measured directly. For PMe3, η3propargyl complex 1-t-Bu was not directly observable, and a limit of ∆G° e -5.5 kcal · mol-1 was placed on the stability of the metallacycle. The stability of the metallacycles follows the same order as the nucleophilicity of the phosphines. Steric effects are also important in determining the stability of the metallacycles, as shown by the fact that the metallacycle formed by addition of PPh3 to Me-substituted η3-propargyl complex 1-Me is much more kinetically stable than the metallacycle 11 formed by addition of PPh3 to t-Bu-substituted 1-t-Bu. Kinetics of Exchange and Rearrangement. The kinetics of exchange of phosphines with the metallacycles proceed by a kinetically first-order dissociative mechanism and were measured only for metallacycles 7 and 9. The barrier for exchange of PMe3 with PPh2Me-substituted metallacycle 9 was 5.5 kcal lower than the barrier for PMe3 exchange with deuterium-labeled metallacycle 7-d9. This activation energy difference parallels the thermodynamic stability of the metallacycles and the leaving group ability of the phosphines. Crude measurements of the rates of rearrangement of all the metallacycles to η2-alkyne complexes were made for all the metallacycles. The barriers for rearrangement also paralleled the thermodynamic stability of the metallacycles and the leaving group ability of the metallacycles [∆Gq for PMe3 (25.6) > PPh2Me (20.3) > PPh3 (∼18) > P(C6H4-p-F)3 (∼17)]. Regioselectivity of metallacyclobutene formation was too high to be measured directly for all of the phosphines studied. The regioselectivity of PMe3 addition to 1-t-Bu was carefully determined by an indirect method: comparison of the rate of exchange of PMe3 with deuterium-labeled metallacycle 7-d9 with the rate of rearrangement of metallacycle 7 to η2-alkyne complex 8 (Figure 7). The regioselectivity for addition of PMe3 to the central carbon of 1-t-Bu was 480:1 (∆∆Gq ) 3.2 kcal · mol-1). Similarly, a similar regioselectivity (∆∆Gq ) 3.4 kcal · mol-1) for addition of PPh2Me to the central carbon of 1-t-Bu was estimated from the rate of exchange of PMe3 with rhenacyclobutene 9 and from the rate of rearrangement of 9 to 10. DFT computational studies of rhenium η3-propargyl complexes were undertaken to better understand the regioselective addition of nucleophiles to the central propargyl carbon.

Figure 7. Free energy diagram for reaction of PMe3 with 1-t-Bu.

Discussions of nucleophilic addition to η3-propargyl complexes usually involve comparisons with well-studied η3-allyl complexes. Transition metal η3-allyl complexes usually undergo nucleophilic addition to a terminal carbon,23 although an increasing number of cases of attack at the central carbon have been reported. The most direct evidence for attack at the central carbon of η3-allyl complexes comes from isolation of metallacyclobutanes.24 Formation of cyclopropanes in the reaction of nucleophiles with η3-allyl complexes is best explained by attack at the central carbon followed by reductive elimination (Scheme 10).25 η3-Allyl complexes having a leaving group such as Cl or OR at the central carbon sometimes react with nucleophiles to give products in which the substituent at the central carbon is replaced by a nucleophile and another nucleophile adds to the (23) (a) Trost, B. M.; Verhoeven, T. R. In ComprehensiVe Organometallic Chemistry; Wilkinson, G., Stone, F. G. A., Abel, E. W., Eds.;Pergamon: Oxford, UK, 1982; Vol. 8, Chapter 57. (b) Tsuji, J. Tetrahedron 1986, 42, 4361. (c) Trost, B. M.; Van Vranken, D. L. Chem. ReV. 1996, 96, 395. (d) Trost, B. M.; Crawley, M. L. Chem. ReV. 2003, 103, 2921. (e) Trost, B. M. J. Org. Chem. 2004, 69, 5813. (24) (a) Ephritikhine, M.; Green, M. L. H.; Mackenzie, R. E. J. Chem. Soc., Chem. Commun. 1976, 619. (b) Ephritikhine, M.; Francis, B. R.; Green, M. L. H.; Mackenzie, R. E.; Smith, M. J. J. Chem. Soc., Dalton Trans. 1977, 1131. (c) Adam, G. J. A.; Davies, S. G.; Ford, K. A.; Ephritikhine, M.; Todd, P. F.; Green, M. L. H. J. Mol. Catal. 1980, 8, 15. (d) Periana, R. A.; Bergman, R. G. J. Am. Chem. Soc. 1984, 106, 7272. (e) McGhee, W. D.; Bergman, R. G. J. Am. Chem. Soc. 1985, 107, 3388. (f) Periana, R. A.; Bergman, R. G. J. Am. Chem. Soc. 1986, 108, 7346. (g) Tjaden, E. B.; Stryker, J. M. J. Am. Chem. Soc. 1990, 112, 6420. (h) Wakefield, J. B.; Stryker, J. M. J. Am. Chem. Soc. 1991, 113, 7057. (i) Tjaden, E. B.; Stryker, J. M. Organometallics 1992, 11, 16. (j) Tjaden, E. B.; Schwiebert, K. E.; Stryker, J. M. J. Am. Chem. Soc. 1992, 114, 1100. (k) Schwiebert, K. E.; Stryker, J. M. Organometallics 1993, 12, 600. (l) Tjaden, E. B.; Stryker, J. M. J. Am. Chem. Soc. 1993, 115, 2083. (m) Review of metallacyclobutenes: Jennings, P. W.; Johnson, L. L. Chem. ReV. 1994, 94, 2241. (25) (a) Hegedus, L. S.; Darlington, W. H.; Russell, C. E. J. Org. Chem. 1980, 45, 5193. (b) Carfagna, C.; Maniani, L.; Musco, A.; Sallese, G.; Santi, R. J. Org. Chem. 1991, 56, 3924. (c) Carfagna, C.; Galarini, R.; Musco, A.; Santi, R. Organometallics 1991, 10, 3956. (d) Hoffmann, H. M. R.; Otte, A. R.; Wilde, A. Angew. Chem., Int. Ed. Engl. 1992, 31, 234. (e) Wilde, A.; Otte, A. R.; Hoffmann, H. M. R. J. Chem. Soc., Chem. Commun. 1993, 615. (f) Formica, M.; Musco, A.; Pontellini, R.; Linn, K.; Mealli, C. J. Organomet. Chem. 1993, 448, C6. (g) Carfagna, C.; Galarini, R.; Linn, K.; Lo´pez, J. A.; Mealli, C.; Musco, A. Organometallics 1993, 12, 3019. (h) Hoffmann, H. M. R.; Otte, A. R.; Wilde, A.; Menzer, S.; Williams, D. J. Angew. Chem., Int. Ed. Engl. 1995, 34, 100. (i) Satake, A.; Nakata, T. J. Am. Chem. Soc. 1998, 120, 10391. (j) Satake, A.; Koshino, H.; Nakata, T. Chem. Lett. 1999, 49. (k) Tjaden, E. B.; Stryker, J. M. J. Am. Chem. Soc. 1990, 112, 6420. (l) Grigg, R.; Korders, M. Eur. J. Org. Chem. 2001, 4, 707.

130 Organometallics, Vol. 28, No. 1, 2009

Casey et al.

Scheme 10

Scheme 11

terminal carbon (Scheme 11).26 These reactions have been suggested to proceed by an unusual addition of a nucleophile to the central carbon followed by elimination of the leaving group from the central carbon to give a substituted η3-allyl complex, which is then attacked at a terminal carbon by a second nucleophile. However, an alternative mechanism involves elimination of HX from the η3-allyl complex to produce an η3propargyl complex, which then undergoes “normal” addition of a nucleophile to the central carbon to produce a metallacyclobutene; protonation of the metallacyclobutene can then produce an η3-allyl complex, which is then attacked at a terminal carbon by a second nucleophile. Several computational studies of η3-allyl compounds have addressed the regioselectivity issue and provided insight into the importance of ligand substitution, conformational preference, and the question of orbital control versus charge control.27 Most studies have focused on η3-allyl palladium and platinum complexes.28 Green has reported extended Hu¨ckel molecular orbital calculations (EHMO) on the neutral η3-propargyl molybdenum complex C5H5(η2-HCtCH)Mo(η3-CH2CtCH).29 Although the regiochemistry of nucleophilic addition was not commented upon, the EHMO results indicated a significant difference in calculated atomic charges between the center (positive) and terminal (negative) carbons, which might form the basis of a charge-controlled explanation of addition of nucleophiles to the (26) (a) Ohe, K.; Matsuda, H.; Morimoto, T.; Ogoshi, S.; Chatani, N.; Murai, S. J. Am. Chem. Soc. 1994, 116, 4125. (b) Castan˜o, A. M.; Aranyos, A.; Szabo´, K. J.; Ba¨ckvall, J.-E. Angew. Chem., Int. Ed. Engl. 1995, 34, 2551. (c) Tsai, F.-Y.; Chen, H.-W.; Chen, J.-T.; Lee, G.-H.; Wang, Y. Organometallics 1997, 16, 822. (d) Aranyos, A.; Szabo´, K. J.; Castan˜o, A. M.; Ba¨ckvall, J.-E. Organometallics 1997, 16, 1058. (e) Organ, M. G.; Miller, M. Tetrahedron Lett. 1997, 38, 8181. (f) Organ, M. G.; Miller, M.; Konstantinou, Z. J. Am. Chem. Soc. 1998, 120, 9283. (g) Kadota, J.; Komori, S.; Fukumoto, Y.; Murai, S. J. Org. Chem. 1999, 64, 7523. (h) Kadota, J.; Katsuragi, H.; Fukumoto, Y.; Murai, S. Organometallics 2000, 19, 979. (i) Organ, M. G.; Arvanitis, E. A.; Hynes, S. J. J. Org. Chem. 2003, 68, 3918. (27) (a) Suzuki, T.; Fujimoto, H. Inorg. Chem. 1999, 38, 370. (b) Sakaki, S.; Takeuchi, K.; Sugimoto, M.; Kurosawa, H. Organometallics 1997, 16, 2995. (c) Aranyos, A.; Szabo´, K. J.; Castan˜o, A. M.; Ba¨ckvall, J.-E. Organometallics 1997, 16, 1058. (d) Ward, T. R. Organometallics 1996, 15, 2836. (e) Carfagna, C.; Galarini, R.; Linn, K.; Lopez, J. A.; Mealli, C.; Musco, A. Organometallics 1993, 12, 3019. (f) Curtis, M. D.; Eisenstein, O. Organometallics 1984, 3, 887. (g) Davies, S. G.; Green, M. L. H.; Mingos, D. M. P. Tetrahedron 1978, 34, 3047. (28) (a) Kollmar, M.; Goldfuss, B.; Reggelin, M.; Rominger, F.; Helmchen, G. Chem.-Eur. J. 2001, 7, 4913. (b) Delbecq, F.; Lapouge, C. Organometallics 2000, 19, 2716. (c) Branchadell, V.; Moreno-Man˜as, M.; Pajuelo, F.; Pleixats, R. Organometallics 1999, 18, 4934. (d) Biswas, B.; Sugimoto, M.; Sakaki, S. Organometallics 1999, 18, 4015. (e) Hagelin, H.; Åkermark, B.; Norrby, P. O. Chem.-Eur. J. 1999, 5, 902. (29) Carfagna, C.; Deeth, R. J.; Green, M.; Mahon, M. F.; McInnes, J. M.; Pellegrini, S.; Woolhouse, C. B. J. Chem. Soc., Dalton Trans. 1995, 3975.

Figure 8. HOMO, LUMO, and LUMO+1 of [C5H5(CO)2Re(η3CH2CtCH)]+ (B).

center carbon. Wojcicki and Bursten reported30 calculations of the model η3-propargyl complex [(PH3)2Pt(η3-CH2CtCH)]+ using the Fenske-Hall (FH) molecular orbital method and later DFT calculations31 and concluded that nucleophilic addition to the central η3-propargyl carbon was charge controlled. Since no analogous theoretical studies of mid-transition metal η3-propargyl compounds were available, we carried out DFT calculations on the model complex [C5H5(CO)2Re(η3CH2CtCH)]+ (B) using B3LYP/LANL2DZ,32 available in the Gaussian 9433 and Gaussian 9834 packages. The molecular orbitals of the η3-propargyl fragment may be described as the simple combination of π-systems for its allyl and ethylene fragments. DFT calculations gave an accurate picture of the frontier molecular orbitals (FMOs) for B and provided insight into the regioselectivity of nucleophilic additions. The energy diagram and the molecular orbitals of the HOMO, LUMO, and LUMO+1 of B are shown in Figure 8. If kinetic addition of nucleophiles to B were orbital controlled, the LUMO of B should have a significant coefficient on the central carbon. Since the LUMO calculated for B has a node at the center carbon and is concentrated on the terminal carbons, the observed nucleophilic attack at the central carbon is not orbital controlled. If nucleophilic attack at the η3-fragment were charge controlled, (30) Graham, J. P.; Wojcicki, A.; Bursten, B. E. Organometallics 1999, 18, 837. (31) Hall, M. B.; Fenske, R. F. Inorg. Chem. 1972, 11, 768. (32) Density functional theory (DFT) has been established as an excellent method for calculations involving organometallic compounds. (a) Ziegler, T. Chem. ReV 1991, 91, 651. (b) Ziegler, T. Can. J. Chem. 1995, 73, 743. (33) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Gill, P. M. W.; Johnson, B. G.; Robb, M. A.; Cheeseman, J. R.; Keith, T.; Petersson, G. A.; Montgomery, J. A.; Raghavachari, K.; Al-Laham, M. A.; Zakrzewski, V. G.; Ortiz, J. V.; Foresman, J. B.; Cioslowski, J.; Stefanov, B. B.; Nanayakkara, A.; Challacombe, M.; Peng, C. Y.; Ayala, P. Y.; Chen, W.; Wong, M. W.; Adres, J. L.; Replogle, E. S.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Binkley, J. S.; Defrees, D. J.; Baker, J.; Stewart, J. P.; Head-Gordon, M.; Gonzalez, C.; Pople, J. A. Gaussian 94, ReVision E.2; Gaussian, Inc.: Pittsburgh, PA, 1995.

Nucleophilic Addition to η3-Propargyl Re Complexes

Organometallics, Vol. 28, No. 1, 2009 131

had a node at the central carbon, and that nucleophilic attack at the central carbon involved interaction with LUMO+1. Their conclusion that regioselective attack at the central carbon resulted from a combination of charge control and the availability of a low-lying orbital centered on the middle propargylic carbon is similar to that reported here.

Experimental Section

Figure 9. Walsh orbital diagram for approach of PH3 to the center carbon of B.

the charge on the central carbon of the η3-propargyl unit should be more positive than that on the terminal carbons. The natural charges computed by the natural population analysis35subset of the DFT calculation show a greater positive charge on the center carbon and support the hypothesis that the observed kinetic addition to the central carbon is charge controlled (Figure 8).36 The attack of a nucleophile at the central carbon of an η3propargyl complex can be envisioned as attack at LUMO+1, which has a major contribution from the central carbon. Thus, an incoming nucleophile could interact with the slightly higher energy LUMO+1 orbital to avoid steric repulsions at the C1 and C3 positions, resulting in a preference for central attack. A linear synchronous transit calculation (LST) was performed to approximate addition of the model nucleophile PH3 to the central propargyl carbon of B to give the model metallacyclobutene complex [C5H5(CO)2Re(CH2C(PH3)dCH)]+. Along the reaction coordinate, a rapid decrease in energy of the HOMO (lone pair of PH3) was observed as the lone pair develops a bonding interaction, and an inversion of the LUMO and LUMO+1 orbitals was observed as the nucleophile approaches (Figure 9). This implies that while charge controls the site of nucleophilic attack, the accessible energy and nodal properties of LUMO+1 allow the attack to proceed smoothly. An exclusive LUMO orbital control argument can be made for the observed η2-alkyne and η2-allene rearrangement products arising from nucleophilic attack at the terminal carbons. A similar DFT study of [(PH3)2Pd(η3-CH3CHCtCH)]+ was performed by Delbecq, Sinou, and co-workers.37 Similar to the results reported here for a rhenium cation, they found that the central propargylic carbon was most positive, that the LUMO (34) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov, B. B.; Liu, G.; Liashenko, A.; PiskorzP.; Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian 98, ReVision A.9; Gaussian, Inc.: Pittsburgh, PA, 1998. (35) (a) Glendening, E. D.; Badenhoop, J. K.; Reed, A. E.; Carpenter, J. E.; Weinhold, F. NBO 4.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, 1996. (b) Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735. (36) The Mulliken charges show a similar trend, with the central carbon being most positive: CH2(+.091)sC(+0.357)sCH(+0.005). (37) Labrosse, J.-R.; Lhoste, P.; Delbecq, F.; Sinou, D. Eur. J. Org. Chem. 2003, 2813.

Rate of Rearrangement of 7 to 8. A CD3CN solution of 7 (26 mg, 0.04 mmol, 0.085 M) containing toluene as an internal integration standard (10 µL, 8.5 µmol) was monitored by 1H NMR spectroscopy at 64.5 °C over 1.5 h. A plot of ln [7] versus time was linear to over 3 half-lives (R2 > 0.98) and gave krearr ) 7.25 × 10-4 s-1, t1/2 ) 16 min, and ∆Gq ) 24.7 kcal · mol-1. Measurements were repeated on similarly prepared samples of 7 over the temperature range 35-55 °C, and the resulting rates of rearrangement were used in an Eyring plot of kobs versus 1/T: kobs(35 °C) ) 9.2 × 10-6 s-1; kobs(44 °C) ) 3.1 × 10-5 s-1; kobs(54 °C) ) 1.7 × 10-4 s-1 [see Figure 1-S in the Supporting Information for Eyring plot]. These data provided ∆Hq ) 30.1 ( 0.9 kcal · mol-1, ∆Sq ) 16.0 ( 0.9 eu, and were used to calculate ∆Gq10 °C ) 25.6 kcal · mol-1, krearr ) 1.02 × 10-7 s-1, t1/2 ≈ 78 days. [C5Me5(CO)2Re(CH2C[P(CD3)3]dCCMe3)][BF4] (7-d9). An excess of P(CD3)3 was condensed onto a frozen CH2Cl2 solution of 1-t-Bu (28.2 mg, 0.05 mmol) at 77 K. A pale yellow-green solution was obtained upon brief mixing at room temperature. Volatile material was evaporated to give 7-d9 in quantitative yield as an off-white flaky powder. 1H NMR (CD3CN, 360 MHz, -20 °C): δ 0.086 (d, J ) 12.2 Hz, CHH), 1.01 (s, CMe3), 1.40 (dd, J ) 12.3, 1.1 Hz, CHH), 1.85 (s, C5Me5). Rate of Exchange of P(CH3)3 with 7-d9. Excess P(CH3)3 was condensed onto a frozen (77 K) CD3CN solution of 7-d9 (30 mg, 0.047 mmol, 0.107 M) containing toluene (10 µL, 8.5 µmol) as an internal standard for NMR integration. The sample was melted, thoroughly mixed at -78 °C, and transferred to a NMR spectrometer probe precooled to -20 °C. Integration showed that its initial concentration of P(CH3)3 was 0.55 M (5 equiv relative to 7-d9). The NMR probe was warmed to 10 °C (thermocouple calibration), and the appearance of the P(CH3)3 1H NMR resonance of 7 (δ 1.79) was monitored versus toluene internal standard. The reaction was monitored for 1.5 halflives, with a late data point acquired after 3 half-lives [see Figure 2-S in the Supporting Information for kinetic plot]. Exponential fit of the concentration versus time data (R2 > 0.98) gave a calculated equilibrium concentration for 7 (0.0979 M) and kexch ) (4.9 ( 0.5) × 10-5 s-1, t1/2 ) 3.9 h, and ∆Gq ) 22.1 kcal · mol-1.

Acknowledgment. Financial support was provided by the National Science Foundation (CHE-0209476). Grants from the NSF (CHE-9629688) and NIH (I S10 RR04981-01) for the purchase of NMR spectrometers are acknowledged. We thank Dr. Douglas R. Powell for assistance with X-ray crystallography, Dr. Charles Fry for assistance with NMR spectrometry, and Professor Frank A. Weinhold for assistance with computations. Supporting Information Available: General experimental information, experimental procedures, additional free energy diagrams for reactions of phosphines with 1-t-Bu3, information on rearrangement of diphosphine-substituted metallacyclobutenes, selected computational output for [C5H5(CO)2Re(η3-CH2CtCH)]+ (B), X-ray crystallographic data for 8-Cl and {C5Me5(CO)2Re[η2(Ar2PCH2CH2PPh2)CH2CtCCMe3]}[BF4] · 2CH2Cl2 (18-CH2Cl2). This material is available free of charge via the Internet at http://pubs.acs.org. OM800739J